open access sensors · correlation radiometers xavier bosch-lluis *, isaac ramos-perez, adriano...

22
Sensors 2011, 11, 6066-6087; doi:10.3390/s110606066 sensors ISSN 1424-8220 www.mdpi.com/journal/sensors Article A General Analysis of the Impact of Digitization in Microwave Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab, Signal Theory and Communications Department, Building D3, Universitat Politècnica de Catalunya and IEEC CRAE/UPC, E-08034 Barcelona, Spain; E-Mails: [email protected] (I.R.-P.); [email protected] (A.C.); [email protected] (N.R.-A.); [email protected] (E.V.); [email protected] (H.P.) * Author to whom correspondence should be addressed; E-Mail: [email protected]; Tel.: +34-934-017-362. Received: 18 April 2011; in revised form: 24 May 2011 / Accepted: 2 June 2011 / Published: 3 June 2011 Abstract: This study provides a general framework to analyze the effects on correlation radiometers of a generic quantization scheme and sampling process. It reviews, unifies and expands several previous works that focused on these effects separately. In addition, it provides a general theoretical background that allows analyzing any digitization scheme including any number of quantization levels, irregular quantization steps, gain compression, clipping, jitter and skew effects of the sampling period. Keywords: microwave; correlation; radiometers; digital; sampling; quantization 1. Introduction Microwave radiometry is today a mature technology that was first used in radio-astronomy in the 1930s [1]. Since then, a large number of microwave radiometers have been developed for remote sensing applications to measure a wide range of natural phenomena (for examples, see [2-5]). Continuous technological evolution has given these systems new capabilities and features. One of the most relevant new technologies in the 1960s [6,7] was the digitization of the signals and the capabilities that emerge from specialized processing platforms such as today powerful an omnipresent Digital Signal Processors (DSP) and Field Programmable Gate Arrays (FPGA). Although digitization OPEN ACCESS

Upload: others

Post on 12-Apr-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11, 6066-6087; doi:10.3390/s110606066

sensors ISSN 1424-8220

www.mdpi.com/journal/sensors

Article

A General Analysis of the Impact of Digitization in Microwave

Correlation Radiometers

Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez,

Enric Valencia and Hyuk Park

Remote Sensing Lab, Signal Theory and Communications Department, Building D3, Universitat

Politècnica de Catalunya and IEEC CRAE/UPC, E-08034 Barcelona, Spain;

E-Mails: [email protected] (I.R.-P.); [email protected] (A.C.);

[email protected] (N.R.-A.); [email protected] (E.V.); [email protected] (H.P.)

* Author to whom correspondence should be addressed; E-Mail: [email protected];

Tel.: +34-934-017-362.

Received: 18 April 2011; in revised form: 24 May 2011 / Accepted: 2 June 2011 /

Published: 3 June 2011

Abstract: This study provides a general framework to analyze the effects on correlation

radiometers of a generic quantization scheme and sampling process. It reviews, unifies and

expands several previous works that focused on these effects separately. In addition, it

provides a general theoretical background that allows analyzing any digitization scheme

including any number of quantization levels, irregular quantization steps, gain

compression, clipping, jitter and skew effects of the sampling period.

Keywords: microwave; correlation; radiometers; digital; sampling; quantization

1. Introduction

Microwave radiometry is today a mature technology that was first used in radio-astronomy in

the 1930s [1]. Since then, a large number of microwave radiometers have been developed for remote

sensing applications to measure a wide range of natural phenomena (for examples, see [2-5]).

Continuous technological evolution has given these systems new capabilities and features. One of the

most relevant new technologies in the 1960s [6,7] was the digitization of the signals and the

capabilities that emerge from specialized processing platforms such as today powerful an omnipresent

Digital Signal Processors (DSP) and Field Programmable Gate Arrays (FPGA). Although digitization

OPEN ACCESS

Page 2: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6067

provides versatility, re-configurability and other advantages for microwave radiometry, it also has side

effects that must be carefully analyzed.

Digitization effects can be separated into quantization and sampling effects. Mainly, the effects

related with the quantization of the input signal (thermal noise) imply the loss of its statistical

properties due to the non-linear quantization process. Consequently, it is not possible to apply the

well-known Gaussian statistical relationships to the quantified signal. This effect has a large impact

when limited quantization levels are considered, and it can be mitigated by increasing the number of

quantization levels. Non-linear effects studies on Gaussian signals started with the early analysis of the

spectrum of clipped noise by Van Vleck and Middleton [8]. In the late 1950s, Price published a work

focusing on the relationship between the ideal correlation between two random signals with Gaussian

probability density function (pdf), and the correlation measured after a non-linear manipulation of

these random signals [9]. This relationship is now used to study the effects of arbitrary quantization

schemes on the correlation of two signals.

Sampling has an impact on the correlation of the two input signals by creating spectral replicas.

Additional noise can be added to the mean correlation value due to spectra replication depending on

the ratio between the sampling frequency and the signal’s bandwidth. Moreover, the sampling period

and its inaccuracies (skew and jitter in the sampling periods) of the Analog to Digital Converter (ADC)

can affect and distort the sampled signal and so the cross-correlation value.

To compare different sampling and quantization schemes an equivalent integration time is defined

as the one required to obtain the same resolution as in the ideal (analog) correlation. Hagen and

Farley [10] conducted important work on this topic, where the effective time was defined and some

easy-to-calculate digitization configurations were analyzed in depth.

More recent works have extended the initial digital radiometer proposed by Weinreb in 1961 [6] for

auto-correlation spectrometers in the radio-astronomy field to include total power radiometers,

polarimeters [11], and digitization impact on interferometric radiometers [12-15].

This work provides a general context to analyze the digitization effects on the cross-correlation of

two Gaussian random signals in the most general case including any number of quantization levels,

irregular quantization steps, gain compression, clipping, bandwidth, sampling frequency, and the skew

and jitter inaccuracies of the sampling periods.

In Section 2 a general analysis of the correlation over non-linear functions applied to Gaussian

noise is given. In Section 3 quantization effects are analyzed as a particular case of the results of the

previous section. Section 4 analyses the sampling process. Section 5 evaluates the correlation variance

due to the digitization. Finally, Section 6 presents the main conclusions of this study.

2. Non-Linearity Impact on the Correlation of Two Gaussian Random Signals

The input signal of a microwave radiometer comes from natural thermal emission. Let us consider

here two independent Gaussian random processes with a certain cross-correlation coefficient

characterized by their probability density function (pdf). Furthermore, both input signals are assumed

to be statistically identical and stationary, so the properties of the signals at a given time (T0) are

independent of T0. Furthermore, both signals fulfill that their statistical properties can be deduced from

a single, sufficiently long sample of the process (realization), i.e., they are ergodic processes [16].

Page 3: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6068

Thus, the variance of each series is constant and the covariance between elements depends only on

their time separation. Finally, let x and y be input signals of a radiometer fulfilling the previous

requirements. Usually, these random variables have zero mean (µx,y = 0), except for offsets, and are

considered to have a standard deviation different from one . In this case, the joint probability of density

is defined as in Equation (1):

(1)

where:

are the standard deviation of x and y signals, and

is the Pearson correlation coefficient and it is defined as the ratio between the covariance

of x and y and the geometric mean of the product of their variances (Equation (2)):

(2)

where the brackets indicate a statistical average, over a time sequence with identical statistics

(ergodic properties). Price’s theorem [9] relates the correlation coefficient of x and y after ( ) and

before ( ) the non-linear functions (

and ) to the random variables (Equation (3)):

(3)

where:

is the correlation coefficient of x and y after the non-linear functions and

are

applied to x and y, respectively

is the ideal correlation coefficient of x and y (before the non-linear functions

and )

are the standard deviation of x and y signals, (for the sake of clarity, from now

on ), and

k denotes the kth

derivative of each function.

From Equation (3), it is possible to retrieve , the ideal correlation between x and y, by

k-derivation and further correlating the and

signals and then k-integrating with respect to

the variable from −1 to +1. These operations can be defined as a relationship between the ideal and

the non-linear correlations. It is defined here as the function

(4)

Considering the correlation of two signals at a delay different form zero (x → x(t = 0) and

y → y(t = )) the effect of the hardware response can be included as depending on , then

Equation (2) becomes:

(5)

Page 4: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6069

where:

is the Pearson correlation coefficient (Equation (2)) at = 0, and

is the so-called Fringe-Washing Function

(FWF), as defined in [17], where Bx and By are the equivalent noise bandwidths, f0 is the

central frequency, and and are the normalized frequency responses of the x

and y channels, respectively. Note that, if the channels’ frequency response are equal, with

rectangular shape of bandwidth B, and centered frequency f0, the fringe-washing function

reduces to the well-known formula .

Equation (5) depends on the correlation coefficient and the hardware features. Usually, to avoid an

underestimation of the correlation coefficient, Equation (4) has to be compensated by the existing

delay between x and y as it follows:

(6)

Assuming that the x and y are stationary random processes (thermal noise radiation) that fulfill the

ergodic property, then Equations (2) and (5) can be calculated in practice using the cross-correlation

technique (multiplication and time averaging).

3. Quantization Impact

This section particularizes the previous analysis to a generic ADC function. The degree of

non-linearity of an ADC function depends on the number of quantization levels and the ADC span

window (VADC = ), going from 2 levels (1 bit) the most non-linear scheme, up to an infinite

number of levels with an infinite VADC, which can be considered linear. As explained before,

quantization does not take into account the sampling, so that this analysis considers an infinite

frequency sampling (Fs → ∞). Sampling effects will be added in Section 5.

Figure 1. Generic analog-to-digital transfer function. The function plotted has compression

gain and different steps in analog and digitized domains.

-1 -0.5 0 0.5 1

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

Analog Domain (x)

Dig

itiz

ed D

om

ain

- f

unc(X

n)

Generic Analog-to-Digital function

Xi-1

Xi

Xi+1

X set

i

X0

XM-1

Page 5: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6070

Equation (7) shows a general quantization function (Figure 1):

(7)

where:

x is the value of the signal in the analog domain (continuous in time and values),

and are two consecutive threshold values of the input signal in the space (see

Figure 1), where is the set of all the input threshold values in the digital domain

( m = 0… ). Note that, in the most general case the distance between two

consecutive steps does not have to be constant ( ,

is the Heaviside function (step function) centered at ,

is the quantization step value between the interval. Note that, in the most

general case the distance between two consecutive steps does not have to be constant either

, and

the values and define the lower and upper bounds of the quantization window.

Again, in a general case, the lower bound can be different from the upper bound

( ).

The first derivative function of Equation (7) is given by:

(8)

where is the Kronecker’s delta: = 1 if and = 0 if .

Substituting the results obtained in Equation (8) in Equation (3), with different functions for

x and y, and considering the first derivative function (k = 1), the relationship between the correlation of

two signals before and after the quantization process is obtained in Equation (9):

(9)

Thereafter, it is easy to compute the integrals over the x and y domains by evaluating them at the

points where Kronecker’s delta does not vanish. Therefore, Equation (9) becomes Equation (10):

(10)

and finally, the integration over the domain is performed from −1 to 1.

(11)

From Equation (11) it is straightforward to include the impulse response of the system by

considering several values of for several delays ( ):

Page 6: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6071

(12)

In some particular and simplified cases, Equation (12) can be obtained analytically, for instance in

the case of quantifying with two levels (one bit, and

). In this case,

Equation (12) becomes the well-known solution stated in Equation (13) [9]:

(13)

Otherwise, when the quantization scheme is more complex Equation (12) can only be analyzed

numerically, such as having different numbers of bits, when signals are clipped, when the quantization

steps are irregular, when the ADC has a non-linear circuit before that exhibits gain compression, when

there is no symmetry between the positive and negative parts or many other possibilities.

Figure 2 has been obtained from Equation (12) for several quantization schemes. Two functions

are plotted for reference, the coarsest quantization scheme (i.e., 2 levels/1 bit, Equation (13)), and the

ideal one (i.e., infinite levels, ). Three more functions are shown considering an ADC

span window of 5σx,y (VADC = 5σx,y), 15 quantization levels or 3.9 bits ( , note that it is not

necessary to have an integer number of bits), but changing the non-linear function. Several

functions have been considered such as equally spaced levels, using gain compression modeled by an

hyperbolic tangent ( ) or having randomly spaced levels. All results are different from

the others, showing the need of a deeper analysis.

Figure 2. Relationship between the non-linear and the ideal correlation for different

digitization schemes.

The root mean square error (RMSE) coefficient between and has been chosen to

measure the degree of linearity of . So, the RMSE has been selected to be the figure of merit for

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

q[·] function for 15 levels correlation for VADC

=5

ideal correlation

non-lin

ear

corr

ela

tion

equally spaced

1bit/2 level

ideal

gain compression

randomly spaced

Page 7: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6072

the comparison of several quantization schemes. Obviously, the linearity of any function is better

as the RMSE is closer to zero. For a given application, there is a maximum distortion of the RMSE that

can be afforded by the and that is called MaxRMSE. If the relationship between the quantized and

the ideal correlations is linear enough (RMSE < MaxRMSE), then the value of the quantized correlation

matches with the ideal correlation and there is no need of using the function. Otherwise, if it is

not linear enough (RMSE ≥ MaxRMSE) then, the obtained correlation has to be modified by the

function in order to retrieve the real correlation.

Figure 3 presents the RMSE computed using Equation (12), with different quantization levels and

VADC. Each curve has a minimum that corresponds to the optimum configuration of VADC with respect

σx,y. As the number of quantization levels increases, the minimum value of RMSE curve gets closer

to 0, and the VADC/σx,y range where the curves remain close to 0 increases as well, since the function

q[·] is linear over a wider range of input powers. By inspecting Figure 3 it is clear that for VADC < 2σx,y,

clipping has a dominant impact over the non-linear correlation. This effect has the maximum impact

when the clipping reduces the whole ADC to a 2 level decision (1 bit). As the clipping effect increases,

the RMSE value converges to the RMSE of the 1 bit/2 levels quantization (RMSE =14.2%), which is

the most non-linear quantization scheme possible.

Figure 3. Root mean square error and ADC span window relationship for different equally

spaced quantification levels.

On the other hand, plots also have a common trend in the region above the minimum, where the

ratio VADC/σx,y increases. As the number of quantization levels increases, the ideal and non-linear

relationship (q[·])becomes more linear over a wider range of input power (the region around the

minimum widens). But when VADC/σx,y →∞ , the RMSE asymptotically increases again up 14.2% (the

value of the 1 bit/2 levels) for all the quantization schemes. This effect can be understood as fewer bits

are being used as the VADC/σx,y ratio increases, and it can be understood as a decreasing of the effective

0 2 4 6 8 10 12 14 16 18 200

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

RMSE for different equally spaced quantization levels

VADC

/x,y

RM

SE

2 Levels

3 Levels

7 Levels

31 Levels

15 Levels

Page 8: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6073

number of bits. The maximum effect occurs when the VADC is so spread over σx,y that only two levels

are effectively used for quantization.

Another conclusion that can be drawn from Figure 3, it is that each quantization level has its own

optimum ADC span window (VADC) with relation of the standard deviation of the input signal. This is

a critical design parameter and it has been summarized on Table 1 for some quantization schemes.

Table 1 shows that the two-level quantization scheme is not sensitive to the VADC/σx,y relationship

and the RSME is always 14.2% and, on the other hand for 31 levels the relationship between quantized

and ideal correlation is almost linear. Furthermore, as the number of levels increases the minimum

RMSE decreases exponentially (Figure 4) and it occurs at a higher VADC/σx,y ratio.

Table 1. Summary of the critical design parameter VADC/σx,y. Relationship between its

optimal configuration and the minimum root mean square error obtained for some

quantization schemes.

Levels VADC/σx,y optimum RMSE minimum [%]

2 (1 bit) no optimum 14.2

3 (1.6 bits) 0.7 4.7

7 (2.8 bits) 1.8 1.4

15 (3.90 bits) 2.2 0.3

31 (4.95 bits) 3.0 0.02

63 (5.97 bits) 3.5 0.0066

Figure 4. Minimum root mean square errors for different quantization levels, it decreases

following an exponential trend, (a) RMSE decreasing vs. the quantization levels, and

(b) RMSE in a semi-log axis decreasing vs. the number of bits.

(a) (b)

Figure 5 shows the effect of using various functions. It compares an equally spaced

quantification with a hyperbolic tangent spacing, used to compress the gain.

0 10 20 30 40 50 60 700

5

10

15

quantization levels

RM

SE

[%

]

Minimum RMSE for different quantization levels

1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 610

-3

10-2

10-1

100

101

102

Number of bits

RM

SE

[%

]

Minimum RMSE for different number of bits

Page 9: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6074

Figure 5. Root mean square error for different levels and different spaced quantification

levels, (a) 7 quantization levels, and (b) 15 quantization levels.

(a) (b)

In both cases, Figure 5(a) (seven levels) and Figure 5(b) (15 levels), there are no changes in the

clipping area, but there are as the VADC/σx,y ratio increases. The effect is that an ADC with gain

compression has lower RMSE values than the same ADC but with equally spaced level scheme for a

wider area. This is equivalent to increasing the number of quantization levels. An ADC with gain

compression can be useful to ensure linearity over a wide range of input signal power, reducing the

necessary number of bits. A common radiometric application where a high range of input power is

measured is during the radiometric Thot − Tcold calibration.

4. Impact of Quantization on the Correlation Spectrum

In [7] some analytically easy-to-solve quantization schemes were analyzed using Taylor series

approximations when <<1, which applies in most radio-astronomical measurements, but not

necessarily in Earth remote sensing or during a radiometric calibration. The cross-correlation

spectrum, the frequency power density is spread and distorted due to quantization step and clipping. In

this work, the spectrum is computed following Equation (12). Therefore, the analysis can be

exhaustive and include different quantization levels for two input signals for any complexity, gain

compression, different bandwidths, and different channel frequency responses. The frequency power

density is the Fourier transformation of the cross-correlation function, which is the Wiener-Khinchin

theorem (Equation (14)):

(14)

Furthermore, Equation (14) also shows the relationship of the cross-correlation spectrum with the

individual signals spectrum, which is the multiplication of the Fourier transforms of the output signals

of the quantization functions. By replacing Equation (4) in Equation (14) the spectral power density of

the non-linear cross-correlation can be readily obtained (Equation (15)):

0 2 4 6 8 10 12 14 16 18 200

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16RMSE for 7 quantization levels

VADC

/x,y

RM

SE

Equally spaced

Gain compression

0 2 4 6 8 10 12 14 16 18 200

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16RMSE for 15 quantization levels

VADC

/x,y

RM

SE

Equally spaced

Gain compression

Page 10: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6075

(15)

where the function can be obtained analytically (following Equation (13)) or numerically

(following Equation (12)).

The effect of the quantization in the cross-correlation spectrum is to decrease and distort it in the

pass-band and spread it over the rejected band. Figure 6 presents an example of the impact of the

quantization on the spectrum computed following Equation (15). The input signals, x and y, that have

been considered for this analysis have a rectangular frequency channel’s response with a pass-band

bandwidth of 2B and a VADC = 5σx,y relationship for all the quantization schemes.

Figure 6(a) presents the x and y linear cross-correlation on the delay domain for several quantization

levels. The lower the number of levels, the higher the correlation distortion. Figure 6(b) shows the

corresponding spectrum of cross-correlations presented in Figure 6(a), which have been computed

following Equation (15). As expected, the two-level scheme has the highest spectrum distortion, the

spectrum spread increases on the rejected band and the pass-band decreases.

Figure 6. Effect of the quantization on the cross-correlation spectrum, (a) ideal

cross-correlation ( ) and cross-correlation for different quantization levels, and

(b) their corresponding spectrum.

(a) (b)

Table 2. Summary of the spectrum distortion and spread for VADC = 5σx,y and different

quantization schemes.

Levels Pass-band Rejected band increase

2 (1 bit) 80% 13%

7 (2.8 bits) 87% 5.0%

15 (3.90 bits) 97% 1.3%

31 (4.95 bits) 99% 1.0%

Ideal 100% 0.0%

-3/(2B) -1/B -1/(2B) 0 1/(2B) 1/B 3/(2B)0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Delay

Norm

aliz

ed F

WF

FWF for VADC

=5x,y

Ideal

2 levels

7 levels

15 levels

31 levels

-4B -3B -2B -B 0 B 2B 3B 4B0

0.2

0.4

0.6

0.8

1

Frequency

Norm

aliz

ed s

pectr

um

FFT(Cross-Correlation) for VADC

=5x,y

Ideal

2 levels

7 levels

15 levels

31 levels

Page 11: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6076

Table 2 summarizes these results, the main conclusion is that using VADC = 5σx,y and at least 15

quantization levels it can be assumed that it neither spreads nor distorts the spectrum. Figure 7 shows

the effect for two functions and for seven quantization levels. It can be observed that the

compression gain reduces the distortion and the spread spectrum.

Figure 7. Effect of the quantization on the cross-correlation spectrum, (a) ideal

cross-correlation ( ) and cross-correlation for two different quantization schemes

using 7 levels, and (b) presents their corresponding spectra.

(a) (b)

In this section, it has been shown that the quantization can be successfully expressed and analyzed

as a non-linear transformation. Despite the quantization can affect significantly the value of the

cross-correlation, it is possible to recover the ideal correlation using the inverse function which

can be obtained numerically. Furthermore, for each quantization scheme there is an optimum

configuration, in terms of linearity, of the VADC/σx,y relationship. Furthermore, quantization distorts and

spreads the cross-correlation spectrum. Finally, the quantization not only has an impact on the value of

the cross-correlation and on the spread of the spectrum, but it increases the standard deviation of the

cross-correlation, as it will be shown in the next section.

5. Sampling Impact

The main effect of sampling is that it creates replicas of the spectrum. Usually, in a standard digital

radiometer topology there is an anti-aliasing filter before the digitization process. Even in this case, the

digitization can introduce aliasing due to the spread of the quantized spectrum, which depends on the

sampling frequency (Fs), and the baseband bandwidth (B). To understand how this effect impacts on

the sampling it is necessary to define the Signal to Noise Ratio (SNR) as in Equation (16):

(16)

where is the expected non-linear cross-correlation value, and is the standard deviation.

-3/(2B) -1/B -1/(2B) 0 1/(2B) 1/B 3/(2B)0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Delay

Norm

aliz

ed F

WF

FWF for VADC

=5x,y

for 7 levels

Ideal

Equally spaced

Gain compression

-4B -3B -2B -B 0 B 2B 3B 4B0

0.2

0.4

0.6

0.8

1

Frequency

Norm

aliz

ed s

pectr

um

FFT(Cross-Correlation) for VADC

=5x,y

for 7 levels

Ideal

Equally spaced

Gain compression

Page 12: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6077

Moreover, the SNR term is related with the radiometric resolution [18], which is one of the figures

of merit of any radiometer. The SNR can decrease due to an increase of the standard deviation of the

correlation or due to a decrease of the , i.e., a noise increment on the pass-band (Equation (2)).

Figure 8 shows an ideal baseband spectrum with an infinite sampling frequency ( , in

black) and a cross-correlation spectrum spread due to quantization ( , in green) for different

Fs and its replicas [19,20]. After quantization and sampling, and further cross-correlation of the

signals, the quantization tail of the Nyquist’s replica (Fs = 2B, in red) overlaps the pass-band of the

baseband replica, so an extra amount of noise is included in the measurement of , and the effect is

a decrease of the calculated mean value (Equation (2)). On the other hand, the standard deviation remains

constant, but the SNR decreases, which has a large impact when considering sub-Nyquist sampling.

Figure 8. Effect of the sampling on the correlation spectrum, (F{·} stands for Fourier

transform). Spectrum of the ideal and the quantized correlation at baseband, black and

green, respectively. In red, it is the first spectrum replica for the minimum sampling

frequency which fulfills the Nyquist criterion (Fs = 2B). In blue, it is the first spectrum

replica for Fs >> 2B.

If the sampling frequency increases, then the first replicas (Fs >> 2B, in blue) appear further away

from the band-pass replica and the tail effect has less impact than before. This is why when only few

quantization levels are used an increase of the sampling frequency has a significant impact on the

radiometric resolution. Even though, it has less impact when more quantization levels are used.

Figure 9 shows the impact on the quantized correlation of different sampling frequencies. The

considerations for these plots are a 2 quantization levels scheme, a noise equivalent band-pass

bandwidth of 2B (assuming a band-pass signal), and VADC = 5σx,y, taking into account the previous

considerations about the spectrum replicas. Figure 9(a) shows the deformation of the fringe washing

function for a set of sampling rates. The sub-Nyquist sampling (Fs = 0.75 FNyquist) has the largest

deformation: it is wider, sharper than the rest of them, and the side lobes are higher.

For the exactly Nyquist sampling rate the fringe-washing function is still sharp and wide. For a

higher sampling rate (Fs > 2 FNyquist) the function is not affected by the sampling rate. On the other

hand, the spectrum in Figure 9(b) is totally distorted in the sub Nyquist sampling rate case, and

recovers its original shape as the sampling rate increases.

Page 13: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6078

Figure 9. Impact on the quantized correlation of different sampling frequencies.

For 2 quantization levels, a bandwidth 2B (assuming a band-pass signal), and VADC = 5σx,y.

(a) shows the fringe-washing function deformation, and (b) shows the effect on their spectrum.

(a)

(b)

Figure 10 shows the impact on the quantized correlation of different sampling frequencies. The

difference from the previous analysis is that in these plots 31 quantization levels have been considered.

As this quantization scheme spread less the spectrum than the previous one, a better performance is

achieved for lower sampling rates. Figure 10(a) shows the distortion of the fringe-washing function

considering different sampling rates. The sub-Nyquist sampling (Fs = 0.75 FNyquist) has the largest

distortion: it is wider, sharper than the rest of them, and the side lobes are higher. For a larger sampling

rate (Fs ≥ FNyquist) the function is not affected by the sampling rate. On the other hand, in the

Figure 10(b), the spectrum is completely distorted for the sub-Nyquist sampling rate case, and recovers

-3/(2B) -1/B -1/(2B) 0 1/(2B) 1/B 3/(2B)0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Fringe-Washing Function for different Fs, [2B/3L and V

ADC/

x,y=5]

Delay [ns]

Norm

aliz

ed F

WF

F

s=0.75F

Nyquist

Fs=F

Nyquist

Fs=2F

Nyquist

Fs=8F

Nyquist

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 10

2

4

6

8

10

12

14

Correlation Spectrum for different Fs, [3L/2b and V

ADC/

x,y=5]

Digital Frequency

Spectr

um

Fs=0.75F

Nyquist

Fs=F

Nyquist

Fs=2F

Nyquist

Fs=8F

Nyquist

Page 14: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6079

its original shape as the sampling rate increases. As expected, it has better performance for higher

sampling rates.

Figure 10. Impact on the quantized correlation of different sampling frequencies.

For 31 quantization levels, a bandwidth 2B, and VADC = 5σx,y. (a) shows the deformation

for the fringe-washing function, and (b) shows the effect on their spectrum.

(a)

(b)

6. Sampling Rate Inaccuracies

In the conversion from analog to digital signals, the sampling frequency is normally assumed to be

constant. Nevertheless, the sampling clock is prone to interferences, thermal drifts and other effects

that can change the oscillating frequency and introduce sampling inaccuracies.

-3/(2B) -1/B -1/(2B) 0 1/(2B) 1/B 3/(2B)0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Fringe-Washing Function for different Fs, [31L/5b and V

ADC/

x,y=5]

Delay

Norm

aliz

ed F

WF

Fs=0.75F

Nyquist

Fs=F

Nyquist

Fs=2F

Nyquist

Fs=8F

Nyquist

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 10

2

4

6

8

10

12

14

16

18

Correlation Spectrum for different Fs, [31L/5b and V

ADC/

x,y=5]

Digital Frequency

Norm

aliz

ed F

WF

Fs=0.75F

Nyquist

Fs=F

Nyquist

Fs=2F

Nyquist

Fs=8F

Nyquist

Page 15: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6080

Figure 11 shows the effect of these inaccuracies, that can be summarized into skew (TSkew), which is

the delay offset between the sampling time of x and y, and jitter, which is the time fluctuation above

the mean value (TJitter). The impact of the skew is changing the point where the FWF is evaluated and a

phase rotation of the correlation coefficient if the working frequency is not exactly zero. Meanwhile,

jitter creates a perturbation for each correlation sample around the mean point with zero mean.

Figure 11. Impact of the clock inaccuracies on the correlation value. Tskew is the sampling

rate offset between x and y, and TJitter is its fluctuations due to clock inaccuracies.

A clock skew error not only has an effect on the modulus of the correlation coefficient, but it affects

its phase by rotating the real and imaginary parts [21], as it is shown in Equation (17):

(17)

where:

, the correlation coefficient is a complex value with modulus

and phase ( )

is the ratio between the residual frequency ( ) and the sampling frequency,

residual digital frequency of the signals, and

is the fringe-washing function evaluated at , and

and are the sampling periods for the x and y inputs, respectively. Both sampling times

are identical except for the skew and jitter effects ( ).

-150 -100 -50 0 50 100 1500

0.2

0.4

0.6

0.8

1

Delay [ns]

Norm

aliz

ed F

WF

Skew and Jitter Analysis

TSkew

FWF correction due to TSkew

Averaged FWF correction due to T

Skew+T

Jitter

TJitter

Max. excursion

Page 16: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6081

On the other hand, the jitter is a random variable effect that can be statistically modeled as a

polynomial of standard deviation depending on time ( ), where n and

m are two different samples. So that, it can be associated to two different and independent

processes ([22,23]):

the aperture jitter, which is related to the random sampling time variations in the ADC

caused by thermal noise in the sample&hold circuit. The aperture jitter is commonly

modeled as an independent Gaussian variable with zero mean and a standard deviation of

, and

the clock jitter, which is a parameter of the clock generator that drives the ADC with the

clock signal. The clock jitter is modeled as a Wiener process, i.e., a continuous-time

non-stationary random process with independent Gaussian increments with a standard

deviation equal to .

Taking into account a wideband signal in jitter terms (fulfill the inequality ). From

here, x can be defined using its discrete-time Fourier transformation (Equation (18)):

(18)

where is the value of the instantaneous jitter. Thence, following Equation (18) the correlation can

be defined again using Equations (14) and (18):

(19)

(20)

where:

is the expected value of the jitter and it depends on the jitter type [23],

is the existing jitter between the mth sample of x and the nth

sample y, and

is the digital delay between the cross-correlation samples ( ).

Equation (21) applies for an aperture jitter and Equation (22) applies for the clock jitter:

and

if n≠m else 1 (21)

(22)

where is the phase noise constant of the oscillator. Thence, the impact of the aperture jitter to the

cross-correlation coefficient is shown in Equation (24):

(23)

Page 17: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6082

(24)

where:

is the cross-correlation coefficient modified by the jitter,

is the cross-correlation coefficient, and

denotes the convolution operator,

From Equation (22) it can be inferred that the aperture jitter does not depend on the sampling rate

and it has a low-pass filter effect on the correlation spectrum with its cut-off frequency depending on

. Particularizing Equation (24) in the case of having a rectangular spectrum, then Equation (24)

becomes Equation (25), and the cross-correlation function is multiplied by an

function

(

):

(25)

As it can be seen, in the case of aperture jitter the coherence loss ratio (Equation (20)) is

independent of the number of sampling points N (therefore of the integration block length TsN), while

in the case of clock jitter it strongly depends on it (Equation (21)).

7. Analysis of the Cross-Correlation Variance due to the Digitization

In the previous sections, quantization and sampling have been analyzed in terms of the impact on

the retrieved value of the correlation. In this section, the variances are analyzed taking into account

both effects (quantization and sampling). In a general case, the output of a correlator of two quantized

and sampled signals after averaging Nq samples is defined as:

(26)

where:

is the sampling period for the x and y inputs, respectively. In most of the

applications both sampling times are identical except for skew and jitter effects

( ), for simplicity, without loss of generality, skew and jitter

can be simulated as well), and

are defined in Equation (3).

Obviously, the product of a pair of samples of in Equation (26)

does not follow a Gaussian pdf. Rather, the distribution of follows a Rayleigh pdf when the g

functions are linear (no quantization). In a more general case, having non-linear functions, the

distribution of follows different trends depending on the quantization scheme. However, the

average of many of these products, over a large number of pairs of samples, approaches a Gaussian

pdf, as implied by the central limit theorem. The variance of the correlator output can then be

calculated as in Equation (27):

Page 18: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6083

(27)

where the first term of this addition is expanded as Equation (28):

(28)

Thence, Equation (28) is split in two parts, the first one takes into account only the elements which

have the same n and m time indexes, and a second part, which is the rest of the Equation (28):

(29)

the first term of Equation (29) can be computed using the properties of the fourth and second statistical

moments relationship for zero mean Gaussian random variables [18], as follows:

(30)

where:

and

are the variances of and , respectively, and

is the modulus of the cross-correlation coefficient of the input signals after the

non-linear function.

The second term in Equation (30) can be expanded using the same statistical moments relationship:

(31)

where:

and are the auto-correlation of the and respectively taking

into account all the samples between sample 1 and the Nq, as in Equation (12).

Page 19: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6084

The first part of the variance of the correlation is obtained by combining the results obtained in

Equations (30) and (31):

(32)

Finally, the value of the variance of the -samples averaged correlation is given by:

(33)

and arranging Equation (33), the variance appears in its usual form, as it is known in the literature [17]:

(34)

where,

is the variance for each sample of the correlation (

).

Following Equation (25), Figure 12 shows the evolution of

with the integration time. The

signals considered in these plots have a noise equivalent squared bandwidth of pass-band of 2B,

, and VADC = 5σx,y.

Figure 12. Impact of the number of averaged samples on the

, (a) shows the impact

for a 3 level quantization scheme and different sampling rate, and (b) shows a constant

sampling rate changing the number of quantization levels.

(a)

20 40 60 80 100 120 140 160 180

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Integration Time [samples]

R

xy

Rxy

Vs integration samples [ VADC

/xy

=5]

3 Level

7 Level

15 levels

Page 20: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6085

Figure 12. Cont.

(b)

Figure 12(a) shows the results for the three quantization levels, all the curves converge to zero

asymptotically. The Fs = 0.75FNyquist curve converges faster, because its samples are uncorrelated

among them. For a given number of averaged samples, the variance of the quantized correlation

increases with the sampling rate, because successive samples are more and more correlated. Recall

that, abscises axis of these plots are displayed in number of averaged samples, not in integration time

units. Obviously, for a given integration time, the plot with the highest sample rate has the

lowest

, it has had time to average more samples despite that for a given number of samples it

exhibits the worst performance.

In Figure 12(b) the sampling rate has been frozen to Fs = FNyquist and two quantization schemes has

been analyzed (3 and 15 levels). The three-level curve converges more quickly to zero than

the 15 levels curve. This is because the fringe washing function is shaper due to the higher non-linear

distortion in three levels than in the 15 levels quantization. This effect is equivalent to have more

uncorrelated samples. Despite the mean of the correlation is lower for the three-level that the 15-level

quantization, the variance of the correlation is also lower in this case.

8. Conclusions

The impact of a general quantization scheme has been expressed and analyzed as a non-linear

transformation. Despite the quantization can significantly affect the value of the ideal cross-correlation

( ), it is possible to recover the ideal correlation ( ) from the measured one ( ) using then the

function which can be always numerically obtained. It has been found that for each

quantization scheme there is an optimum configuration of the VADC/σx,y relationship for which the

linearity with respect the ideal correlation is minimized. Furthermore, quantization distorts and spreads

the cross-correlation spectrum, decreasing the radiometric resolution.

0 200 400 600 800 1000 12000

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Integration Time [samples]

R

xy

Rxy

Vs integration samples [3L/2b VADC

/xy

=5]

3 Levels

15 Levels

7 Levels

Page 21: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6086

Sampling also has an impact in the spectrum of the cross-correlation function, and depending on the

sampling rate, replicas of the spectrum can overlap the main spectrum. This can be due to a sampling

rate below the Nyquist criterion or, even above the Nyquist criterion, due to the non-linear quantization

process which spreads the spectrum. In both cases, there is an impact on the SNR, and so on the

radiometric resolution. Other sampling effects have been analyzed such as the clock inaccuracies:

skew and jitter.

Finally, the impact of quantization and sampling on the variance of the measurements has been

studied. It has been found that it strongly depends on the relationship between the bandwidth of the

cross-correlation and the sampling rate. As the sampling rate increases, the successive samples are

more correlated, so they add less new information to the measurements and the variance

decreases slowly.

Acknowledgments

This work has been conducted as part of the award “Passive Advanced Unit (PAU): A Hybrid

L-band Radiometer, GNSS Reflectometer and IR-Radiometer for Passive Remote Sensing of the

Ocean” made under the European Heads of Research Councils and European Science Foundation

European Young Investigator (EURYI) awards scheme in 2004, it was supported in part by the

Participating Organizations of EURYI, and the EC Sixth Framework Program. It has also been

supported by the Spanish National Research and EU FEDER Project TEC2005-06863-C02-01 (FPI

grant BES-2006-11578, 2006-2010), and by the Spanish National Research and EU FEDER Project

AYA2008-05906-C02-01/ESP.

References

1. Dicke, R.H. The measurement of thermal radiation at microwave frequencies. Rev. Sci. Instr.

1946, 17, 268-275.

2. Le Vine, D.M.; Lagerloef, G.S.E.; Colomb, F.R.; Yueh, S.H.; Pellerano, A. Aquarius: An

instrument to monitor sea surface salinity from space. IEEE Trans. Geosci. Remot. Sens. 2007, 45,

2040-2049.

3. Westwater, E.R. The accuracy of water vapor and cloud liquid determination by dual frequency

ground-based microwave radiometry. Radio Sci. 1978, 32, 677-685.

4. Barre, H.M.J.; Duesmann, B.; Kerr, Y.H. SMOS: The mission and the system. IEEE Trans.

Geosci. Remot. Sens. 2008, 46, 587-593.

5. Aja, B.; Artal, E.; de la Fuente, L.; Pascual, J.P.; Mediavilla, A.; Roddis, N.; Kettle, D.;

Winder, W.F.; Cara, L.P.; de Paco, P. Very low-noise differential radiometer at 30 GHz for the

PLANCK LFI. IEEE Trans. Microw. Theor. Tech. 2005, 53, 2050-2062.

6. Weinreb, S. Digital radiometer. Proc. IEEE 1961, 49, 1099.

7. Cooper, B.F. Correlators with two-bit quantization. Aust. J. Phys. 1970, 23, 521-527.

8. van Vleck, J.H.; Middleton, D. The spectrum of clipped noise. Proc. IEEE 1966, 54, 2-19.

9. Price, R. A useful theorem for nonlinear devices having Gaussian inputs. IRE Trans. Inform.

Theory 1958, 4, 69-72.

Page 22: OPEN ACCESS sensors · Correlation Radiometers Xavier Bosch-Lluis *, Isaac Ramos-Perez, Adriano Camps, Nereida Rodriguez-Alvarez, Enric Valencia and Hyuk Park Remote Sensing Lab,

Sensors 2011, 11

6087

10. Hagen, J.; Farley, D. Digital-correlation techniques in radio science. Radio Sci. 1973, 8,

775-784.

11. Piepmeier, J.R.; Gasiewski, A.J.; Almodovar, J.E. Advances in microwave digital radiometry.

Proc. IEEE Int. Geosci. Remot. Sens. Symp. 2000, 7, 2830-2833.

12. Fischman, M.A.; England, A.W. Sensitivity of a 1.4 GHz direct sampling digital radiometer.

IEEE Trans. Geosci. Remot. Sens. 1999, 37, 2172-2180.

13. Ruf, C.S. Digital correlators for synthetic aperture interferometric radiometry. IEEE Trans.

Geosci. Remot. Sens. 1995, 33, 1222-1229.

14. Fischman, M.A.; England, A.W.; Ruf, C.S. How digital correlation affects the fringe washing

function in L-band aperture synthesis radiometry. IEEE Trans. Geosci. Remot. Sens. 2002, 40,

671-679.

15. Corbella, I.; Torres, F.; Camps, A.; Bara, J.; Duffo, N.; Vall-llosera, M. L-band aperture synthesis

radiometry: Hardware requirements and system performance. Proc. IEEE Int. Geosci. Remot.

Sens. Symp. 2000, 7, 2975-2977.

16. Papoulis, A. Probability, Random Variables, and Stochastic Processes; McGraw-Hill: New York,

NY, USA, 1991; pp. 427-442.

17. Camps, A.; Torres, F.; Bará, J.; Corbella, I.; Monzón, F. Automatic calibration of channels

frequency response in interferometric radiometers. Electron. Lett. 1999, 35, 115-116.

18. Tiuri, M.E. Radio astronomy receivers. IEEE Trans. Antenn. Propag. 1964, 12, 930-938.

19. Thompson, A.R.; Moran, J.M.; Swenson, G.W. Interferometry and Synthesis in Radio Astronomy,

2nd ed.; Wiley-VCH: Berlin, Germany, 2007.

20. Ramos-Perez, I.; Bosch-Lluis, X.; Camps, A.; Rodriguez-Alvarez, N.; Marchan-Hernandez, J.F.;

Valencia, E.; Vernich, C.; de la Rosa, S.; Pantoja, S. Calibration of correlation radiometers using

pseudo-random noise signals. Sensors 2009, 9, 6131-6149.

21. Camps, A.; Torres, F.; Corbella, I.; Bara, J.; Lluch, J.A. Threshold and timing errors of 1 bid2

level digital correlators in earth observation synthetic aperture radiometry. Electron. Lett. 1997,

33, 812-813.

22. Demir, A.; Mehrotra, A.; Roychowdhury, J. Phase noise in oscillators: A unifying theory and

numerical methods for characterization. IEEE Trans. Circ. Syst. Fund. Theor. Appl. 2000, 47,

655-674.

23. Kobayashi, H.; Morimura, M.; Kobayashi, K.; Onaya, Y. Aperture jitter effects in wideband ADC

systems. In Proceedings of the 38th SICE Annual Conference, Iwate, Japan, 28–30 July 1999;

pp. 1089-1094.

© 2011 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article

distributed under the terms and conditions of the Creative Commons Attribution license

(http://creativecommons.org/licenses/by/3.0/).